You are on page 1of 15

energies

Article
Application of Field Olfactometry to Monitor the Odour Impact
of a Municipal Sewage System
Andrzej Kulig , Mirosław Szyłak-Szydłowski * and Marta Wiśniewska

Faculty of Building Services, Hydro and Environmental Engineering, Warsaw University of Technology,
20 Nowowiejska Street, 00-653 Warsaw, Poland; andrzej.kulig@pw.edu.pl (A.K.);
marta.wisniewska@pw.edu.pl (M.W.)
* Correspondence: miroslaw.szydlowski@pw.edu.pl

Abstract: Odorant emissions are associated with, among other things, wastewater transport in sewer
networks; they contribute to air pollution and result in complaints from residents living close to
emission sources. The critical location in terms of the formation of unpleasant odour compounds
is the pressure line that connects the pumping station and the expansion well; this is where they
are released into the atmosphere. This paper presents comprehensive results of olfactometric and
chromatographic tests in the Polish city of Białystok using portable devices that allow for multiple
determinations and instant results. The study attempts to investigate the relationship between odour
and odorant concentrations and check the suitability of field olfactometry as a tool for the ongoing
monitoring of the emission of noxious odours and for verifying complaints submitted by residents.
Statistical analysis shows a very high correlation coefficient between cod and the concentrations
of individual odorants, ranging from 0.82 to 0.91. This olfactometric research, mainly conducted
in situ, can be an appropriate method for the ad hoc monitoring of processes in sewage networks.
Citation: Kulig, A.; Szyłak-
This method allows the detection of unwanted emissions of odours at individual points in the
Szydłowski, M.; Wiśniewska, M.
network in concentrations that are not detected by standard sensors but that nevertheless cause
Application of Field Olfactometry to
odour nuisances, complaints, and social conflict. The research results provide evidence in favour of
Monitor the Odour Impact of a
the energetic usage of wastewater, which is in line with circular economy conception, since odour
Municipal Sewage System. Energies
2022, 15, 4015. https://doi.org/
nuisance is one of its indicators.
10.3390/en15114015
Keywords: environment monitoring; odorants; odour impact; olfactometry; portable GC; sewage system
Academic Editors: Jacek Makinia,
˛
Jan Oleszkiewicz, Zbysław
Dymaczewski, Joanna Jeż-Walkowiak
and J˛edrzej Bylka
1. Introduction
Received: 20 April 2022 1.1. Municipal Sewage Systems as Sources of Odorant Emissions
Accepted: 28 May 2022
In recent years, odour problems caused by sewage, waste, and human and animal
Published: 30 May 2022
activities have become increasingly important [1–7]. Emissions of odorants—the chemical
Publisher’s Note: MDPI stays neutral compounds that cause odours—accompany, among other things, the transport of sewage.
with regard to jurisdictional claims in This contributes to atmospheric pollution as well as a deterioration in air quality and often
published maps and institutional affil- results in complaints from residents living near the sources of the emissions. The most
iations. frequently emitted odorants from installations intended for wastewater transport are hy-
drogen sulphide (H2 S), methanethiol (CH3 SH), and other organic sulphides (e.g., dimethyl
sulphide (DMS) (CH3 )2 S and diphenyl sulphide (C6 H5 )2 S). For example, the concentration
of H2 S in sewers can be up to 314,000 µg/m3 , while the concentration of CH3 SH can reach
Copyright: © 2022 by the authors.
18,300 µg/m3 , and the concentration of DMS can reach 98 µg/m3 . In addition, there is an
Licensee MDPI, Basel, Switzerland.
important compound that is odourless but has explosive properties: methane (CH4 ) [8,9].
This article is an open access article
Both H2 S and CH4 may have negative effects on the environment and also on human
distributed under the terms and
health. H2 S has a sharp, irritating odour, and its olfactory detection threshold is very
conditions of the Creative Commons
low (only 0.0035 ppm [10]); it is, therefore, the most frequent cause of complaints from
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
sewer system users [11,12]. Due to its high toxicity, it is also one of the main causes
4.0/).
of death in the wastewater sector. In addition, H2 S may be responsible for corrosive

Energies 2022, 15, 4015. https://doi.org/10.3390/en15114015 https://www.mdpi.com/journal/energies


Energies 2022, 15, 4015 2 of 15

processes in the sewage infrastructure, causing the degradation of pipes and wells [13].
The production of sulphides in wastewater collection and treatment facilities is particularly
well-known in countries with warm climates, such as Kuwait [14]. In recent years, studies
on the process of microbiologically induced corrosion resulting from the reaction of H2 S
oxidation to sulphuric acid have been conducted by Grengg et al. [15], Nielsen et al. [16],
and Jiang et al. [17]. In these studies, under anaerobic conditions, methanogenesis also
took place during the reduction of sulphates (a chemical process that produces odourless
CH4 with a lower explosion limit equal to approx. 5% by volume and that contributes to
global warming [18]).
The problem of capturing emissions (especially sulphur compounds) concerns both
new and old sewage systems. The most frequent sources of emissions of odorous com-
pounds include sewers, sumps and expansion sewers, pressure pipelines, and sewage
pumping stations [19]. In addition, the odours of household sewage cause a nuisance
to well-being and quality of life, as well as dangerous problems for human health [20].
As a result of oxygen deficiency, wastewater is corrosive, and the anaerobic decomposition
of organic compounds contained in wastewater and sludge deposits occurs. The main
source of sulphur is, in this context, sulphate, with concentrations between 40 mg/dm3 and
200 mg/dm3 . This process involves sulphate-reducing bacteria and methanogenic archaea
that contribute to the degradation of CH3 SH [21,22].
The critical location in terms of the formation of unpleasant odour compounds is
the pressure line that connects the pumping station and the expansion well; this is where
they are released into the atmosphere. Additionally, expansion wells located in the close
vicinity of households increase the problems of odour nuisance. H2 S continuous monitoring
devices are readily available and in widespread use, but however useful they may be for
H2 S, they do not necessarily apply to the rest of the reduced sulphur compounds [23].
The choice of the method used to assess the intensity of odour nuisance often results
from such factors as the variability of sampling conditions and their composition over
time. These factors may influence the results obtained to a greater or lesser degree [24].
Importantly, there is no single generally accepted technique to effectively assess the envi-
ronmental impact of odorous nuisance compounds due to variabilities in hedonic tone and
the chemical character of odorous emissions [25,26].

1.2. Field Olfactometry


One of the methods for quantifying odour emissions is olfactometry. Olfactometry
methods may be categorised according to where the sample is analysed: indirect olfactome-
try (the gas sample is captured in a dedicated bag and then transported to the laboratory
for analysis) and direct olfactometry (the gas sample is analysed at the source of the odour).
In the case of indirect tests, there is a risk that chemical reactions occur during transport
inside the bag and affect the result obtained in the laboratory [27,28].
Field olfactometers were developed to overcome the drawbacks of sample storage
and evaluate low cod in ambient air [29]. Hayes et al., wrote that field testing eliminates
the potential dangers of sample degradation in laboratory-based testing and offers a better
opportunity to analyse different areas at different times but at the same site. In addition,
it greatly reduces the capacity for an odour to degrade, due to testing taking place almost
immediately after sample collection [30]. Recent research has shown that dynamic olfac-
tometry is an alternative technique to monitoring bioprocesses, reducing the number of
analyses and their time and cost [31–36]. Field olfactometry can be used for proactive
monitoring or as an enforcement tool for confident odour measurement at property lines
and at locations throughout a community near emission sources [37]. Olfactometers, which
are simplified portable dilution devices, help to determine odour levels and give a reading
of the D/T (Dilution-to-Threshold) ratio [38–40]—the dilution of an odour sample that
cannot be distinguished from odourless air by 50% of the members of an odour panel.
In addition, olfactometers are a useful tool for downwind odour intensity (iod ) measure-
Energies 2022, 15, 4015 3 of 15

ment [41]. The use of this technique, followed by the application of dispersion models,
permits the quantification of the odour impact of a process [42].
Currently, two main instruments called Nasal Ranger (St. Croix Sensors, Inc.) and
SM-100 (IDES Canada Inc.) are used. Szyłak-Szydłowski found no evidence to reject the
null hypothesis of equal average values of D/T obtained during THT and H2 S research
using the above olfactometers—the achieved results were not significantly different [43].
Other researchers found that the Nasal Ranger performed well in generating dilutions for
all examined compounds and that only at the highest dilution was a discrepancy found
between the set and observed dilution ratios. The Scentroid SM-100 showed a linear
relationship between the set and observed dilution ratios. However, higher observed
dilution ratios (up to a factor of 2) were observed when compared to the dilution ratio
set points [44]. Examples of the use of field olfactometers in various research areas are
summarised in Table 1.

Table 1. Examples of the use of field olfactometers in various research areas.

Field of Study Reference


Impacts of mechanical-biological waste treatment plants and
[33,36]
validating the cod from piles
Assessment of the odour nuisance of railway sleepers saturated
[45]
with creosote oil
Estimation of iod in the atmospheric air near municipal
[46]
wastewater treatment plants (WWTP)
Determination of cod around a water reclamation plant [47]
Determination of cod around a sewage sludge composting plant [48,49]
Assessing the range of olfactory impacts of municipal landfills [49]
Determination of cod near a sewage sludge biodrying installation [50]
Measurement of the effects of WWTP modernisation [51]
Monitoring livestock farm odours [41,52]
Evaluation of odour nuisance coupled with odour dispersion
modelling can be applied to the development of local urban [53,54]
planning strategies
Monitoring odours of dairy manure slurry [55]
Determination of cod around poultry barns [56]

Neither cod nor odorant concentration measurements conducted alone allow for ob-
taining the full range of information necessary to characterise the odour nuisance of a
studied object [28,57]. One of the methods used to assess the level of odour stimulation and
contributions of individual odorants to the odour is the Odour Activity Value (OAV), also re-
ferred to as the Odour Activity Concentration (OAC) [23,35,58]. The parameter is defined
as the concentration of a substance (Ci) to its odour detection threshold (ODT). However,
studies carried out by Gonzales et al. [49], among others, indicate a very weak correlation
between OAV and cod , which is mainly due to the adoption of the ODT value, which
is different in different literature sources; it is also due to the masking of the individual
compounds included in the odour mixture.
Evaluations of environmental quality are challenging to implement in urban and inte-
grated planning, strategic environmental assessments, and other environmental procedures,
given the often non-spatial character of the obtained results [52]. Therefore, it is important
to extend the standard analytical methodologies for monitoring extensive areas, such as
sewage networks, with olfactometric analyses. This paper presents comprehensive results
of olfactometric and chromatographic tests (H2 S, CH3 SH, and (CH3 )2 S—DMS) in the Polish
city of Białystok. The research is essential from the point of view of finding the cause of
numerous complaints about odour nuisance reported by the city’s inhabitants. The study
attempts to find the relationship between the cod and the concentration of the tested odor-
ants and check the usefulness of field olfactometry as a tool for the ongoing monitoring of
the emission of odours and for verifying the complaints reported by residents.
Energies 2022, 15, 4015 4 of 15

The main objective of this work is to implement procedures for the determination
of cod in the assessment of the quality of the urban environment. These procedures can
significantly complement determinations of the concentrations of selected odorants, in-
fluencing the scent concentration value and, as a result, the smell nuisance. A secondary
objective is to investigate the primary relationships between cod and the essential odorous
compounds detected in sewage systems. It may be helpful to introduce the energetic usage
of wastewater in future.

2. Materials and Methods


2.1. Research Schedule
Potential sources of odour emissions were identified, inventoried, and classified in
the north-western part of the city of Białystok, where, in the years 2015–2017, residents
registered complaints about the presence of odour nuisances. The research work included:
• collecting and compiling data on residents’ complaints and potential sources of odour
emissions in the analysed area;
• field tests for odorants;
• the analysis and elaboration of the research results and the formulation of conclusions.
The data on residents’ complaints and potential sources of emissions were supple-
mented and verified during the local inspection in the analysed area and during the field
research. The direct field research on the concentrations of selected odorants and odour
was aimed at ascertaining the air quality conditions inside the sewers, in facilities in the
network, and in atmospheric air. Control of meteorological conditions accompanied the
field tests. Table 2 contains the schedule of tests performed and the number of receptor
points determined for chemical and olfactometric tests.

Table 2. Schedule of chemical and olfactometric tests with the number of receptor points.

Number of Points
Series No. Research Date Chemical Olfactometry
Measurements Measurements
I 14–15 November 2017 19 48
II 16–17 January 2018 32 32
III 13–14 April 2018 20 36
IV 25–26 May 2018 20 55
V 14–15 July 2018 20 86
VI 12 September 2018 20 76
Sum: - 131 333

2.2. Location of the Research Areas


Measurements and observations were carried out in the north-western part of Białys-
tok (Figure 1). Based on survey research, analysis of potential odour sources, and local
complaints, the following areas were identified:
• Area 1: the area containing nodal sewage chambers;
• Area 2: the area containing nuisance sewage manholes in the immediate vicinity of an
apartment block;
• Area 3: the vicinity of the “Fasty” industrial wastewater;
• Area 4: the area along the street with the last section of the main sewer connecting to
the municipal WWTP;
• Other: surroundings of single manholes for sewage collectors in different parts of
the city.
 Area 3: the vicinity of the “Fasty” industrial wastewater;
 Area 4: the area along the street with the last section of the main sewer connecting to
Energies 2022, 15, 4015
the municipal WWTP; 5 of 15
 Other: surroundings of single manholes for sewage collectors in different parts of the
city.

Figure 1. Areas of the research with the specification of key sampling locations.
Figure 1. Areas of the research with the specification of key sampling locations.

2.3. Characteristics of the Municipal Sewage System


The analysed sewage system discharges municipal sewage to the sewage treatment
plant; itit is
plant; is aamixture
mixtureofofdomestic
domesticand andindustrial
industrialsewage.
sewage.InIn the
the sewage
sewage system
system of of
thethe
citycity
of
of Białystok,
Białystok, there
there areare
twotwootherother types
types of networks,
of networks, namely
namely the general
the general sewage
sewage systemsystem
and
the
andstormwater
the stormwater drainage system,
drainage whose
system, integration
whose with the
integration withsanitary network
the sanitary was taken
network was
into
takenaccount
into account to the to
extent necessary
the extent to determine
necessary odour odour
to determine nuisances.
nuisances.
The municipal sewage system operates using a gravity-pressuregravity-pressure system.
system. It is a mixed
system featuring a gravitational network cooperating with intermediate pumping stations
that pump
pump sewage sewagefromfromplaces
placeswith
with lower
lower to higher
to higher elevations,
elevations, usually
usually veryvery far away.
far away. The
The pumped wastewater is directed mainly to expansion wells installed
pumped wastewater is directed mainly to expansion wells installed in the network, from in the network,
from
whichwhich it continues
it continues to by
to flow flow by gravity.
gravity.

2.4. Olfactometric Examinations


2.4. Olfactometric Examinations of
of ccod
od
The
The receptor points were located both
receptor points were located both on
on the
the windward
windward side
side (for
(for the
the determination
determination
of
of air pollution background caused by odorants) and on the leeward side of
air pollution background caused by odorants) and on the leeward side of the
the studied
studied
objects. Their detailed locations considered the current wind direction (WD), among other
objects. Their detailed locations considered the current wind direction (WD), among other
factors. The iod was evaluated in sensory studies according to six stages (i = 0–5).
factors. The iod was evaluated in sensory studies according to six stages (i = 0–5). ®
The olfactometers used were portable field olfactometers such as Nasal Ranger ® and
The olfactometers used were portable field olfactometers such as Nasal Ranger and
Scentroid SM-100. They allowed the creation of a calibrated series of dilutions by mixing
Scentroid SM-100. They allowed the creation of a calibrated series of dilutions by mixing
air contaminated with odorants with filtered air free from odours. Field olfactory nuisance
air contaminated with odorants with filtered air free from odours. Field olfactory nuisance
testing using field olfactometers is a method of quantifying the cod in the “Dilution-to-
testing using field olfactometers is a method of quantifying the cod in the “Dilution-to-
Threshold Ratios” (D/T) system. The D/T value is the number of dilutions needed to make
Threshold Ratios” (D/T) system. The D/T value is the number of dilutions needed to make
the odour undetectable in the ambient air. Thanks to the application of the olfactometers
mentioned above, it is possible to calculate the D/T value, which, in turn, leads to the cod
value determined in units of odour (ou—odour unit, analogous to the European standard
PN-EN 13725:2007) per unit of volume [ou/m3 ] [51].
Energies 2022, 15, 4015 6 of 15

2.5. Chemical Examinations of Odorant Concentrations


The studies on the selected odorants, namelyH2 S, CH3 SH, and DMS ((CH3 )2 S), were
carried out using the Photovac Voyager portable gas chromatograph (GC). This is a portable,
automatic gas analyser for identifying airborne chemicals and measuring their concen-
trations. The Voyager uses a GC to collect and analyse the air samples. It uses a 10.6 eV
photoionization detector (PID), a pre-column, and three columns for heavy (C7–C12), mid-
dle (C3–C7), and light (C1–C3) compounds): 4 m × 0.53 mm × 2.0 um SPB-35 (pre-column),
8 m × 0.25 mm BLANK Fused Silica (column A), 20 m × 0.32 mm × 1.0 um Supelcowax10
(PEG) (column B), and 15 m × 0.32 mm × 12 um Quadrex 007–1 (column C). The carrier
gas used for determinations was high-purity nitrogen. The column oven in the Photovac
Voyager is isothermal, 55 to 80 ◦ C. The Voyager can be effectively used to monitor many
of the volatile organic compounds (VOCs) listed in EPA Method 8240A, including chlori-
nated and aromatic hydrocarbons. Method detection limits for VOCs range from parts per
trillion (ppt) in water (ng/dm3 ) to about 500 parts per million (ppm) in the ambient air,
depending upon the type of compound and the detector used [59]. The limit of detection
for the selected odorants was 0.001 ppm, while the limit of quantitation was 0.005 ppm [60].
In addition, analyses of CH3 SH and DMS were performed, but concentrations of those
compounds were lower than the limit of detection.

2.6. Meteorological Examinations


The air tests were accompanied by an assessment of the current meteorological condi-
tions (WD, wind speed (WS), air temperature (AT), relative humidity (RH), and degree of
cloudiness (DC)). The information on meteorological conditions was used directly during
the tests and recorded for every measurement as data necessary for the analysis of cod .
The meteorological parameters were determined at each of the receptor points. WD was
determined by the streak method every time before the beginning of observations and
measurements. WS was measured with a hand anemometer, the Kestrel 4500 NV, with a
wing rotor. The WD and WS were measured at a height of 2 m. AT and RH were measured
at a height of 1.5 m. The Rotronic HygroPalm psychrometer with a HygroClip2 HC2-S3
sensor was used for the measurements. The DC was determined as the degree of cloud
cover in the sky. To determine the DC, an octant scale from 0 to 8 was used (8 means full
cloud cover, while 0 indicates none).

2.7. Statistical Analysis


One-way analysis of variance was used to compare means across populations. The to-
tal variance (variation in outcomes) was divided into a portion derived from the differences
between populations (treatments) and a portion derived from the differences between out-
comes within populations (random error). The analysis of variance only gives information
about whether there are statistically significant differences between populations.
In addition, the value of the Pearson correlation coefficient was calculated. It was
located in the closed interval [−1, 1]. The greater its absolute value, the stronger the linear
relationship between the variables. The linear correlation coefficient can be thought of as a
normalised covariance.
Basic regression parameters, a method based on linear combinations of variables and
parameters that fit the model to the data, were also calculated. The fitted regression line or
curve represents the estimated expected value of variable Y at specific values of another
variable or variables X.

3. Results and Discussion


3.1. Sampling Air from under and above the Manhole to the Sewer
At selected test points, samples were taken from the canal manholes at various depths
and at the level of the canal manholes. Table 3 presents the values of minimum, average,
median, and maximum cod results.
Energies 2022, 15, 4015 7 of 15

Table 3. Values of minimum, average, median, and maximum results of odour and selected odorant
concentrations at test points at different depths from under the manholes to the sewers. Depths:
2.5 m, 1 m, at the level of the manhole, and at the height of 1.8 m.

Sampling Level
Parameter Value
−2.5 m −1 m 0m 1.8 m
min 515 94 0 0
mean 2487 1245 160 17
cod (ou/m3 )
median 1500 750 44 6
max 10,000 10,000 750 106
min 0.020 0 0 0
H2 S concentration mean 1.852 0.524 0.042 0
(mg/m3 ) median 1.144 0 0 0
max 8.715 5.584 0.274 0
min 0 0 0 0
CH3 SH concentration mean 0.166 0.009 0 0
(mg/m3 ) median 0.015 0 0 0
max 1.720 0.093 0 0
min 0 0 0 0
(CH3 )2 S
mean 0.059 0.003 0 0
(DMS) concentration
median 0 0 0 0
(mg/m3 )
max 0.942 0.065 0 0

H2 S was between 0 and 8.7 mg/m3 , depending on sample depth, and Sivret et al.,
reported a median value between 1.9 and 16.0 mg/m3 , depending on the sewer lo-
calisation [61], so the achieved results fit the “standard” values of H2 S concentrations
in sewer networks. Austigard et al., found high values of H2 S in a sewer network:
56.0 mg/m3 downstream and 354 mg/m3 upstream [62]. Mantos et al., measured values
between 145 mg/m3 and 482 mg/m3 [63], whereas Zhang et al., reported H2 S concen-
trations up to 300 mg/m3 [64]. However, these values were observed in closed pipes,
whereas the present study was conducted by sampling through manhole covers from
which H2 S was emitted into the outside air. CH3 SH and DMS values were compa-
rable to those reported by Sivret: the medians, respectively, were 0.3–4.3 mg/m3 and
0.06–0.45 mg/m3 [65]. Hwang et al., measured an average CH3 SH concentration in the in-
fluent of a WWTP of about 3–200 times higher than other VOCs, such as DMS, CH3 S2 CH3 ,
and CS2 [66]. Wang et al., as part of a monitoring program for sewers located at 18 dif-
ferent sites in Sydney and Melbourne, concluded that, in both cities, the CH3 SH concen-
tration (0.675–1.421 mg/m3 ) was substantially higher than the concentrations of DMS,
CH3 S2 CH3 , CH3 S3 CH3 , and CS2 (0.008 mg/m3 –0.094 mg/m3 ) [67]. Devai and DeLaune
measured 8.0–8.7 mg/m3 of CH3 SH and 3.8–26.4 mg/m3 DMS in WWTPs in the United
States [68], while Chan and Hanaeus measured 13.7 mg/m3 and 114.1 mg/m3 , respec-
tively, in Swedish WTTPs [69]. Lasardi et al., detected H2 S concentrations in Greek WWTP
sewers between 0.001 and 111.5 mg/m3 , while the CH3 SH concentration was less than
0.466 mg/m3 [70]. Sun et al., measured H2 S concentrations in sewers between 5.575 mg/m3
and 250.9 mg/m3 [21].
Additionally, a one-factor analysis of variance was performed, examining the signifi-
cance of the influence of the grouping factor, i.e., the sampling level, on the values of cod
and the concentrations of individual odorants. A graphical interpretation of the results
of the analysis of variance is shown in Figure 2. Both the graphical interpretation and the
value of the F value (the ratio of two mean square values), at p < 0.05, allow the rejection
of the zero hypothesis regarding the lack of significant differences between the obtained
results of dependent variables.
Additionally, a one-factor analysis of variance was performed, examining the
significance of the influence of the grouping factor, i.e., the sampling level, on the
values of cod and the concentrations of individual odorants. A graphical interpreta-
tion of the results of the analysis of variance is shown in Figure 2. Both the graphical
Energies 2022, 15, 4015 interpretation and the value of the F value (the ratio of two mean square values), at
8 of 15
p < 0.05, allow the rejection of the zero hypothesis regarding the lack of significant
differences between the obtained results of dependent variables.

Figure 2. Graphical interpretation of the results of the variable variance analysis: cod (ou/m3 ),
Figure 2. Graphical interpretation of the results of the variable variance analysis: c od (ou/m3),
concentrations of selected odorants (mg/m3 ) against the grouping factor, i.e., sampling level. Effective
concentrations of selected odorants (mg/m 3) against the grouping factor, i.e., sampling level.
hypothesis decomposition; vertical bars denote 0.95 confidence intervals.
Effective hypothesis decomposition; vertical bars denote 0.95 confidence intervals.
Then, for each of the variables, a second factor was determined, revealing the extent to
Then, for each of the variables, a second factor was determined, revealing the
which knowledge of the grouping factor explains the variability of the dependent variable.
Itextent to whichofknowledge
is a proportion of theexplained
the total variance groupingbyfactor explains the
the experimental variability
effect of the
(level). For the
dependent variable. It is a proportion of the total variance explained by the
variables cod , H2 S, CH3 SH, and (CH3 )2 S (DMS), the values of this coefficient—and, at the experi-
mental
same effect
time, the(level). For the
percentage variables
of variance ofcthe
od, H2S, CH3 SH, and (CH3) 2S (DMS), the val-
results of individual variables, which can
be explained by the level—were 33%, 33%,time,
ues of this coefficient—and, at the same 13%, the
andpercentage of variance
10%, respectively. Thisof the re-
measure
sults of individual variables, which can be explained by the level—were
describes the percentage of variance explained by a given effect only for a given sample. 33%, 33%,
13%, and 10%,
Additionally, therespectively. This measure
Pearson correlation describes
coefficients between thethe
percentage
examined of variance
variables ex-
were
plained by a given effect only for a given sample. Additionally, the Pearson
calculated. Table 4 presents the values of these coefficients and basic regression parameters. corre-
lationIn coefficients between
most cases, the the examined
correlation coefficientvariables were
value is very calculated.
high, ranging Table 4 presents
from 0.82 to 0.91.
the values of these coefficients and basic regression parameters.
Smaller values were observed for variable pairs: cod —DMS and H2 S—DMS for levels of
−2.5 m and −1 m and CH3 SH—DMS for a level of −1 m. This correlation is significant—the
coefficient p was lower than 0.05 in each case. By analysing the parameters of the regression
function, it was proved that the measure of adjustment of the regression line to the values
observed is also high in most cases. The highest values of the determination factor were
obtained for the cod —H2 S, cod —CH3 SH, and H2 S—CH3 SH variable pairs, both at the
−2.5 m and −1 m levels. Similar results were achieved by Perez et al., and Gostelow et al.;
for the different types of points in the sewer network, the regression coefficients between
odour and H2 S concentration were 0.91–0.99 [71,72]. A study by Thistlethwayte and
Goleb [73] showed close relationships between the concentrations of sulphides, amines,
aldehydes, and H2 S. The authors argue that although the H2 S test is not sufficient to
determine the level of odour nuisance, it can be taken as an indicator of this nuisance.
Research conducted by authors in dry weather conditions in sewers showed H2 S levels in
the range of 0.2–14 mg/m3 . Investigations by Van Gemert [74] showed H2 S values in the
sewage system between 0.003 and 0.59 mg/m3 .
Energies 2022, 15, 4015 9 of 15

Table 4. Correlation coefficients and basic regression parameters of variables: cod and concentrations
of individual odorants.

Factor R R2 t Constant Y Slope Y Constant X Slope X


Level −2.5 −1.0 −2.5 −1.0 −2.5 −1.0 −2.5 −1.0 −2.5 −1.0 −2.5 −1.0 −2.5 −1.0
Cod
0.91 0.97 0.83 0.94 19.74 34.72 −23.98 −349.5 0.72 0.58 456.7 649.4 1.16 1.61
H2 S
Cod
0.82 0.83 0.67 0.69 12.54 12.88 −178.3 −6.29 0.13 0.01 1758 909.1 5.07 72.53
CH3 SH
Cod
0.54 0.73 0.29 0.53 5.60 9.26 −37.29 −2.95 0.04 0.00 2138 1068 7.82 132.1
DMS
H2 S
0.84 0.80 0.71 0.65 13.59 11.71 −152.4 −0.12 0.17 0.02 1171 181.1 4.10 42.23
CH3 SH
H2 S
0.63 0.76 0.40 0.57 7.21 9.99 −43.43 −0.54 0.06 0.01 1422 258.5 7.29 82.16
DMS
CH3 SH
0.88 0.63 0.77 0.40 16.04 7.00 −3.02 0.64 0.37 0.30 44.47 3.71 2.06 1.30
DMS
Abbreviations: Level—sampling level (m), R—Pearson correlation coefficient, R2 —coefficient of determination,
t—value of statistics t examining the significance of the correlation coefficient, Constant—free linear regression of
Y to X and X to Y, Slope—linear regression coefficient of Y to X and X to Y, respectively.

The results presented vary widely, confirming the need to monitor sewage network
emissions. The highest value (10,000 ou/m3 ) of the cod at depths of 2.5 m and 1 m were
recorded at points located in Area 3 (2 m to the north-east of the “Fasty” industrial wastewa-
ter). Values of cod equal to 7500 ou/m3 were found at one of the Other measurement points
(at a depth of 2.5 m). A concentration of 7500 ou/m3 was also found at a point located in
an uncovered pipe above the chamber in Area 1 (at a depth of 1 m). As for other extremes,
the following values of cod at a depth of 1 m were recorded at the following points:
6000 ou/m3 —in Area 1 and at one of the Other sites;
4300 ou/m3 —in Area 2;
3750 ou/m3 —in Area 3 and at one of the Other sites;
3330 ou/m3 —in Area 1;
2700 ou/m3 —in Area 1 and Area 4 (near the municipal WWTP).
Perez et al., using a stationary olfactometer (samples were transported to the labo-
ratory), also measured maximum values of cod lower than 10,000 ou/m3 in manholes in
summer, except at one point where the cod was 35,000 ou/m3 . That value was related to the
fact that the points selected for the site were ‘critical’ points: the chimney and the manhole
were both located close to the exit of a forced main [72].

3.2. Atmospheric Examinations—Sampling at a Height of 1.80 m


cod values in the atmosphere are presented in Figure 3, while Figure 4 shows the iod
values in the north-western part of Białystok.
As a result of the direct field studies in particular areas of the city, it can be con-
cluded that:
• No other sources of odorant emissions were identified except for those from the
municipal sewage system.
• Sewerage wells are a significant source of emissions of catch compounds (including
H2 S, even in concentrations greater than 100 µg/m3 ), regardless of the season and
weather conditions. The odour nuisance occurred in Area 4 and in other places in
Białystok, such as in Area 1 and Area 2.
• Area 1 is also an essential source of odour impacts. Odour emitters include both
manholes above the sewer and above the canal, which was used not only for rainwater
but also for sewage.
Perez
Perez et
et al.,
al., using
using aa stationary
stationary olfactometer
olfactometer (samples
(samples werewere transported
transported to to the
the
laboratory), also
laboratory), also measured
measured maximum
maximum values
values of
of ccod lower than
od lower than 10,000
10,000 ou/m
ou/m3 in
3 in man-
man-
holes in
holes in summer,
summer, except
except atat one
one point
point where
where the
the ccod was 35,000
od was 35,000 ou/m
ou/m33.. That
That value
value
Energies 2022, 15, 4015 was related to the fact that the points selected for the site were ‘critical’ points:
was related to the fact that the points selected for the site were ‘critical’ points: of 10 the
the
15
chimney and the manhole were both located close to the exit of a forced
chimney and the manhole were both located close to the exit of a forced main [72]. main [72].

3.2.
• Atmospheric
3.2. Atmospheric
Significant odour Examinations—Sampling
Examinations—Sampling at
at a
nuisances were found duringa Height
Height of
of 1.80
1.80 mm of the sewer cham-
the rehabilitation
cbersvalues in
(series in
cod values
od the
no.the atmosphere
V).atmosphere are
Uncontrolledare presented
floodplains in
of raw
presented Figure 3, while
while Figure
sewage3,affecting
in Figure 4
4 shows
the surrounding
Figure shows
the iiod
the values
areas
od informed.
werein
values the north-western
the north-western part
part of
of Białystok.
Białystok.

Figure
Figure 3.
3. ccod inthe
od in
od in theatmosphere
the atmosphereof
atmosphere of the
of the north-western
the north-western part
north-western part of
part of Białystok
of Białystok (at
Białystok (at a
(at aa height
height of
height of 1.8
of 1.8 m).
1.8 m).
m).

Figure 4.
Figure 4. iod in
in the
the atmosphere
atmosphere of
of the
the north-western
north-western part
part of
of Białystok
Białystok (at
(at aa height
height of
of 1.8
1.8 m).
m).
Figure 4. iiod
od in the atmosphere of the north-western part of Białystok (at a height of 1.8 m).

H2 S, even in small concentrations, is not only a nuisance but also poses a danger to
human life and health, especially for those involved in the internal operation of sewage
systems. Therefore, water and sewage companies have an obligation to implement the
necessary safety measures before workers enter the sewage system. There is a role in
these measures for portable measuring devices that make it possible to check the con-
Energies 2022, 15, 4015 11 of 15

centration values of toxic and explosive gases present in the sewer system, such as H2 S
and CH4 . Therefore, the current H2 S in situ field monitoring technologies are inadequate
because of lower detection limits and/or complexity [20,75]. The human nose is more
sensitive to H2 S than most of the analytical methods: the odour threshold of that com-
pound is about 0.47 ppbv in air [76] and between 0.025 and 0.25 µg/dm3 in clean water [77].
This is the concentration at which 50% of humans can detect the odour [20]. A synthetic,
multidisciplinary approach, combining field olfactometry with in situ chemical analysis,
is therefore appropriate.

4. Conclusions
Based on olfactometric and chemical determinations, the following corrective actions
were proposed in order to minimise the odour nuisance of the sources identified in the
sewage network operating in the analysed area of the city:
• The improvement of the operation works in the sewage network through the pur-
chase and possible replacement of specialised equipment designed for and dedicated
exclusively to operation works in sewage networks.
• Maintaining proper functioning of the sewer network in terms of hydraulics, while min-
imising the negative effects on the environment, requires the implementation of on-line
monitoring of filling and flow rates in the sewers. This enables the identification of
irregularities in network operations, e.g., very low levels of filling and filling speeds
(which do take place in the examined sewer network), which favour the development
of anaerobic processes in sewers. On the other hand, high filling levels and overfilling
of the network are also recorded. These may result from, among other things, uncon-
trolled and even illegal discharges of wastewater into the city’s network. Monitoring
should also cover the rainwater network, which makes it possible to record events
involving the illegal discharge of sanitary sewage into the rainwater network.
• The modernisation of connection-and-branch chambers located in large sewers, where
unpleasant odours are found to be a nuisance to the environment, in order to stabilise
the flow without reversing it and while ensuring the proper ventilation of the sewers.
• A full inventorying and securing of the non-drainage tanks in the city area with the
systematic control of their proper use and operation.
• The elimination of rainwater discharge into the sanitary sewage system.
Therefore, olfactometric research combined with the determination of odorants re-
sulted not only in making a “map” of the olfactory impact of the sanitary sewage system
over a wide area but also in the detection of irregularities in the operation of the network
and an analysis of the basic dependence of the odour-odorant relationship at several depths
under the sewer wells. Additionally, these studies have allowed the identification of the
above-mentioned corrective actions.
Increasing urban growth and lifestyle expectations have led to increased public com-
plaints regarding odours from sewer infrastructure, mainly from those in close proximity
to a discharge source such as a manhole, sewer vent, or sewer pumping station. Closing
sewer vents to prevent odour complaints can lead to an increase in humidity and H2 S
accumulation within the sewer, leading to increased biogenic corrosion [78]. Studies about
odours/odorants in sewer networks are mainly focused on sulphuric compounds inside
the pipes of closed sewer networks. This paper provided an improved methodology for
odour/odorant monitoring in sewers, mainly in the vicinity of sewage manholes that
are often the source of odour complaints. A significant added value of the work is the
analysis of the concentration of odours and selected odorants at several depths under
the sewage manhole, at the level of the manhole, and at the height of 1.8, which showed
the stratification of the above-mentioned pollutants and the potential impact on reported
odour nuisances. Olfactometric research, especially when conducted in situ, may be an
appropriate method for the ad hoc monitoring of the processes taking place in sewage net-
works, both as a standalone method and in combination with the methods currently used.
It is particularly important in this context to know the relationship between commonly
Energies 2022, 15, 4015 12 of 15

tested odorants and odour emissions. Olfactometric examination allows the detection
of unwanted emissions of odours at individual points in the network in concentrations
that are not detected by standard sensors but which nevertheless cause odour nuisances,
complaints, and social conflict.

Author Contributions: Conceptualisation, M.S.-S. and M.W.; methodology, A.K., M.W. and M.S.-S.;
investigation, A.K., M.S.-S. and M.W.; data curation, M.S.-S. and M.W.; writing—original draft
preparation, M.S.-S. and M.W.; writing—review and editing, A.K.; visualisation, M.S.-S.; supervision,
A.K.; project administration, A.K.; funding acquisition, A.K. All authors have read and agreed to the
published version of the manuscript.
Funding: This research was funded by the Warsaw University of Technology. IDUB Open Science Programme.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Data available on request.
Conflicts of Interest: The authors declare no conflict of interest.

Abbreviations

AT air temperature
cod odour concentration
DC degree of cloudiness
DMS (CH3 )2 S—dimethyl sulphide
D/T Dilution-to-Threshold Ratio
EPA The United States Environmental Protection Agency
GC gas chromatograph
iod odour intensity
PID photoionization detector
RH relative humidity
VOCs volatile organic compounds
WD wind direction
WS wind speed
WWTP wastewater treatment plant

References
1. Koe, L.C.C. Characterization and removal of odors from refuse material. Environ. Monit. Assess. 1988, 10, 75–84. [CrossRef] [PubMed]
2. Filipy, J.; Rumburg, B.; Mount, G.; Westberg, H.; Lamb, B. Identification and quantification of volatile organic compounds from a
dairy. Atmos. Environ. 2006, 40, 1480–1494. [CrossRef]
3. Le Leuch, L.M.; Subrenat, A.; Le Cloirec, P. Removal of target odorous molecules on to activated carbon cloths. Water Sci. Technol.
2004, 50, 193–198. [CrossRef] [PubMed]
4. Lillo-Ródenas, M.A.; Ros, A.; Fuente, E.; Montes-Morán, M.A.; Martin, M.J.; Linares-Solano, A. Further insights into the activation
process of sewage sludge-based precursors by alkaline hydroxides. Chem. Eng. J. 2008, 142, 168–174. [CrossRef]
5. Muezzinoglu, A. A study of volatile organic sulfur emissions causing urban odors. Chemosphere 2003, 51, 245–252. [CrossRef]
6. Vega, E.; Sánchez-Reyna, G.; Mora-Perdomo, V.; Iglesias, G.S.; Arriaga, J.L.; Limón-Sánchez, T.; Escalona-Segura, S.;
Gonzalez-Avalos, E. Air quality assessment in a highly industrialized area of Mexico: Concentrations and sources of volatile
organic compounds. Fuel 2011, 90, 3509–3520. [CrossRef]
7. Wiśniewska, M.; Kulig, A.; Lelicińska-Serafin, K. Odour Emissions of Municipal Waste Biogas Plants—Impact of Technological
Factors, Air Temperature and Humidity. Appl. Sci. 2020, 10, 1093. [CrossRef]
8. Matos, R.V.; Matias, N.; Ferreira, F.; Matos, J.S. Dynamic Modeling of Hydrogen Sulfide within Enclosed Environments in
Biosolids Recovery Facilities. Water Environ. Res. 2017, 88, 2209–2218. [CrossRef]
9. Pochwat, K.; Kida, M.; Ziembowicz, S.; Koszelnik, P. Odours in Sewerage—A Description of Emissions and of Technical
Abatement Measures. Environments 2019, 6, 89. [CrossRef]
10. Mathews, E.R.; Barnett, D.; Petrovski, S.; Franks, A.E. Reviewing microbial electrical systems and bacteriophage biocontrol as
targeted novel treatments for reducing hydrogen sulfide emissions in urban sewer systems. Rev. Environ. Sci. Bio/Technol. 2018,
17, 749–764. [CrossRef]
Energies 2022, 15, 4015 13 of 15

11. Grengg, C.; Mittermayr, F.; Baldermann, A.; Böttcher, M.E.; Leis, A.; Koraimann, G.; Grunert, P.; Dietzel, M. Microbiologically
induced concrete corrosion: A case study from a combined sewer network. Cem. Concr. Res. 2015, 77, 16–25. [CrossRef]
12. Yongsiri, C.; Vollertsen, J.; Hvitved-Jacobsen, T. Hydrogen sulfide emission in sewer networks: A two-phase modeling approach
to the sulfur cycle. Water Sci. Technol. 2004, 50, 161–168. [CrossRef] [PubMed]
13. Vollertsen, J.; Nielsen, A.H.; Jensen, H.S.; Wium-Andersen, T.; Hvitved-Jacobsen, T. Corrosion of concrete sewers—The kinetics of
hydrogen sulfide oxidation. Sci. Total Environ. 2008, 394, 162–170. [CrossRef] [PubMed]
14. Tomar, M.; Abdullah, T.H.A.; Abdullah, J.H.A. Effect of aeration on generation and emission of hydrogen sulfide in wet wells of
lifting stations. Water Air Soil Pollut. 1995, 81, 385–399. [CrossRef]
15. Grengg, C.; Müller, B.; Staudinger, C.; Mittermayr, F.; Breininger, J.; Ungerböck, B.; Borisov, S.M.; Mayr, T.; Dietzel, M.
High-resolution optical pH imaging of concrete exposed to chemically corrosive environments. Cem. Concr. Res. 2019, 116,
231–237. [CrossRef]
16. Nielsen, K.A.; Tattersall, D.B.; Jones, P.R.; Møller, B.L. Metabolon formation in dhurrin biosynthesis. Phytochemistry 2008, 69,
88–98. [CrossRef]
17. Jiang, L.; Liu, H.; Chu, H.; Zhu, C.; Xiong, C.; You, L.; Xu, J.; Zhang, Y.; Qin, Y. Influence of compression fatigue on chloride
threshold value for the corrosion of steels in simulated concrete pore. Constr. Build. Mater. 2014, 73, 699–704. [CrossRef]
18. Ren, Y.G.; Wang, J.H.; Li, H.F.; Zhang, J.; Qi, P.Y.; Hu, Z. Nitrous oxide and methane emissions from different treatment processes
in full-scale municipal wastewater treatment plants. Environ. Technol. 2013, 34, 2917–2927. [CrossRef]
19. Turkmen, M.; Dentel, S.K.; Chiu, P.C.; Hepner, S. Analysis of sulfur and nitrogen odorants using solid-phase microextraction and
GC-MS. Water Sci. Technol. 2004, 50, 115–120. [CrossRef]
20. Chen, D.; Szostak, P. Factor analysis of H2 S emission at a wastewater lift station: A case study. Environ. Monit. Assess. 2013, 185,
3551–3560. [CrossRef]
21. Sun, J.; Hu, S.; Sharma, K.R.; Ni, B.-J.; Yuan, Z. Degradation of methanethiol in anaerobic sewers and its correlation with
methanogenic activities. Water Res. 2015, 69, 80–89. [CrossRef] [PubMed]
22. Sun, J.; Hu, S.; Sharma, K.R.; Ni, B.-J.; Yuan, Z. Stratified Microbial Structure and Activity in Sulfide- and Methane-Producing
Anaerobic Sewer Biofilms. Appl. Environ. Microbiol. 2014, 80, 7042–7052. [CrossRef] [PubMed]
23. Vitko, T.G.; Suffet, I.H. Corrective factors applied to reduced sulfur compounds in wastewater foul air. Water Environ. Res. 2021,
93, 1487–1495. [CrossRef] [PubMed]
24. Gebicki, J.; Byliński, H.; Namieśnik, J. Measurement techniques for assessing the olfactory impact of municipal sewage treatment
plants. Environ. Monit. Assess. 2016, 188, 32. [CrossRef] [PubMed]
25. Naddeo, V.; Belgiorno, V.; Zarra, T. Odour Impact Assessment Handbook; John Wiley & Sons: Hoboken, NJ, USA, 2012; ISBN 111996928X.
26. Muñoz, R.; Sivret, E.C.; Parcsi, G.; Lebrero, R.; Wang, X.; Suffet, I.H.; Stuetz, R.M. Monitoring techniques for odour abatement
assessment. Water Res. 2010, 44, 5129–5149. [CrossRef]
27. Maurer, D.L.; Bragdon, A.M.; Short, B.C.; Heekwon, A.; Koziel, J.A. Improving environmental odor measurements:comparison of
lab-based standard method and portable odor measurement technology. Arch. Environ. Prot. 2018, 44, 100–107. [CrossRef]
28. Maurer, D.L.; Bragdon, A.M.; Short, B.C.; Heekwon, A.; Koziel, J.A.; Wiśniewska, M. Methods of Assessing Odour Emissions
from Biogas Plants Processing Municipal Waste. J. Ecol. Eng. 2020, 21, 140–147. [CrossRef]
29. Giungato, P.; Di Gilio, A.; Palmisani, J.; Marzocca, A.; Mazzone, A.; Brattoli, M.; Giua, R.; De Gennaro, G. Synergistic approaches
for odor active compounds monitoring and identification: State of the art, integration, limits and potentialities of analytical and
sensorial techniques. TrAC Trends Anal. Chem. 2018, 107, 116–129. [CrossRef]
30. Hayes, J.E.; Stevenson, R.J.; Stuetz, R.M. The impact of malodour on communities: A review of assessment techniques. Sci. Total
Environ. 2014, 500, 395–407. [CrossRef]
31. González, I.; Robledo-Mahón, T.; Silva-Castro, G.A.; Rodríguez-Calvo, A.; Gutiérrez, M.C.; Martín, M.Á.; Chica, A.F.; Calvo, C.
Evolution of the composting process with semi-permeable film technology at industrial scale. J. Clean. Prod. 2016, 115,
245–254. [CrossRef]
32. Gutiérrez, M.C.; Martín, M.A.; Chica, A.F. Usual variables and odour concentration to evaluate composting process and odour
impact. Environ. Technol. 2014, 35, 709–718. [CrossRef] [PubMed]
33. Kulig, A.; Szylak-Szydlowski, M. Assessment of range of olfactory impact of plant to mechanical-biological treatment of municipal
waste. Chem. Eng. Trans. 2016, 54, 247–252.
34. Sironi, S.; Capelli, L.; Céntola, P.; Del Rosso, R.; Pierucci, S. Odour impact assessment by means of dynamic olfactometry,
dispersion modelling and social participation. Atmos. Environ. 2010, 44, 354–360. [CrossRef]
35. Bax, C.; Sironi, S.; Capelli, L. How Can Odors Be Measured? An Overview of Methods and Their Applications. Atmosphere 2020,
11, 92. [CrossRef]
36. Szyłak-Szydłowski, M. Validation of odor concentration from mechanical-biological treatment piles using static chamber and
wind tunnel with different wind speed values. J. Air Waste Manag. Assoc. 2017, 67, 1046–1054. [CrossRef]
37. Nicell, J.A. Assessment and regulation of odour impacts. Atmos. Environ. 2009, 43, 196–206. [CrossRef]
38. Brandt, R.C.; Adviento-Borbe, M.A.A.; Elliott, H.A.; Wheeler, E.F. Protocols for Reliable Field Olfactometry Odor Evaluations.
Appl. Eng. Agric. 2011, 27, 457–466. [CrossRef]
39. Benzo, M.; Mantovani, A.; Pittarello, A. Measurement of Odour Concentration of Immissions using a New Field Olfactometer
and Markers’ Chemical Analysis. Chem. Eng. Trans. 2012, 30, 103–108. [CrossRef]
Energies 2022, 15, 4015 14 of 15

40. Capelli, L.; Sironi, S.; Del Rosso, R.; Guillot, J.-M. Measuring odours in the environment vs. dispersion modelling: A review.
Atmos. Environ. 2013, 79, 731–743. [CrossRef]
41. Pan, L.; Yang, S.X.; DeBruyn, J. Factor Analysis of Downwind Odours from Livestock Farms. Biosyst. Eng. 2007, 96,
387–397. [CrossRef]
42. Toledo, M.; Siles, J.A.; Gutiérrez, M.C.; Martín, M.A. Monitoring of the composting process of different agroindustrial waste:
Influence of the operational variables on the odorous impact. Waste Manag. 2018, 76, 266–274. [CrossRef] [PubMed]
43. Szyłak-Szydłowski, M. Comparison of two types of field olfactometers for assessing odours in laboratory and field tests. Chem.
Eng. Trans. 2014, 40, 67–72.
44. Walgraeve, C.; Van Huffel, K.; Bruneel, J.; Van Langenhove, H. Evaluation of the performance of field olfactometers by selected
ion flow tube mass spectrometry. Biosyst. Eng. 2015, 137, 84–94. [CrossRef]
45. Szylak-Szydlowski, M. Odour nuisance of railway sleepers saturated with creosote oil. Chem. Eng. Trans. 2016, 54, 163–168.
46. Lewkowska, P.; Cieślik, B.; Dymerski, T.; Konieczka, P.; Namieśnik, J. Characteristics of odors emitted from municipal wastewater
treatment plant and methods for their identification and deodorization techniques. Environ. Res. 2016, 151, 573–586. [CrossRef]
47. Wang, T.; Sattayatewa, C.; Venkatesan, D.; Noll, K.E.; Pagilla, K.R.; Moschandreas, D.J. Modeling indoor odor–odorant con-
centrations and the relative humidity effect on odor perception at a water reclamation plant. Atmos. Environ. 2011, 45,
7235–7239. [CrossRef]
48. Pearson, C.; Littlewood, E.; Douglas, P.; Robertson, S.; Gant, T.W.; Hansell, A.L. Exposures and health outcomes in relation to
bioaerosol emissions from composting facilities: A systematic review of occupational and community studies. J. Toxicol. Environ.
Health. B Crit. Rev. 2015, 18, 43–69. [CrossRef]
49. González, D.; Colón, J.; Sánchez, A.; Gabriel, D. A systematic study on the VOCs characterization and odour emissions in a
full-scale sewage sludge composting plant. J. Hazard. Mater. 2019, 373, 733–740. [CrossRef]
50. González, D.; Guerra, N.; Colón, J.; Gabriel, D.; Ponsá, S.; Sánchez, A. Filling in sewage sludge biodrying gaps: Greenhouse gases,
volatile organic compounds and odour emissions. Bioresour. Technol. 2019, 291, 121857. [CrossRef]
51. Kulig, A.; Szyłak-Szydłowski, M. Assessment of the Effects of Wastewater Treatment Plant Modernization by Means of the Field
Olfactometry Method. Water 2019, 11, 2367. [CrossRef]
52. Man, Z.; Dai, X.; Rong, L.; Kong, X.; Ying, S.; Xin, Y.; Liu, D. Evaluation of storage bags for odour sampling from intensive pig
production measured by proton-transfer-reaction mass-spectrometry. Biosyst. Eng. 2019, 189, 48–59. [CrossRef]
53. Badach, J.; Kolasińska, P.; Paciorek, M.; Wojnowski, W.; Dymerski, T.; G˛ebicki, G.; Dymnicka, M.; Namieśnik, J. A
case study of odour nuisance evaluation in the context of integrated urban planning. J. Environ. Manag. 2018, 213,
417–424. [CrossRef] [PubMed]
54. Invernizzi, M.; Capelli, L.; Sironi, S. Proposal of Odor Nuisance Index as Urban Planning Tool. Chem. Senses 2017, 42, 105–110.
[CrossRef] [PubMed]
55. Brandt, R.C.; Elliott, H.A.; Adviento-Borbe, M.A.A.; Wheeler, E.F.; Kleinman, P.J.A.; Beegle, D.B. Field Olfactometry Assessment
of Dairy Manure Land Application Methods. J. Environ. Qual. 2011, 40, 431–437. [CrossRef] [PubMed]
56. Damuchali, A.M.; Guo, H. Evaluation of a field olfactometer in odour concentration measurement. Biosyst. Eng. 2019, 187,
239–246. [CrossRef]
57. Grzelka, A.; Sówka, I.; Miller, U. Metody oceny emisji odorów z obiektów gospodarki hodowlanej. Inżynieria Ekol. 2018, 19,
56–64. [CrossRef]
58. Fisher, R.M.; Barczak, R.J.; Suffet, I.H.M.; Hayes, J.E.; Stuetz, R.M. Framework for the use of odour wheels to manage odours
throughout wastewater biosolids processing. Sci. Total Environ. 2018, 634, 214–223. [CrossRef]
59. Szyłak-Szydłowski, M. Odour Samples Degradation During Detention in Tedlar® Bags. Water Air Soil Pollut. 2015,
226, 227. [CrossRef]
60. Wiśniewska, M.; Kulig, A.; Lelicińska-Serafin, K. Olfactometric testing as a method for assessing odour nuisance of biogas plants
processing municipal waste. Arch. Environ. Prot. 2020, 46, 60–68. [CrossRef]
61. Sivret, E.C.; Le-Minh, N.; Wang, B.; Wang, X.; Stuetz, R.M. Dynamics of Volatile Sulfur Compounds and Volatile Organic
Compounds in Sewer Headspace Air. J. Environ. Eng. 2017, 143, 04016080. [CrossRef]
62. Austigard, A.D.; Svendsen, K.; Heldal, K.K. Hydrogen sulphide exposure in waste water treatment. J. Occup. Med. Toxicol. 2018,
13, 10. [CrossRef] [PubMed]
63. Matos, R.V.; Ferreira, F.; Gil, C.; Matos, J.S. Understanding the effect of ventilation, intermittent pumping and seasonality in
hydrogen sulfide and methane concentrations in a coastal sewerage system. Environ. Sci. Pollut. Res. 2019, 26, 3404–3414.
[CrossRef] [PubMed]
64. Zhang, L.; De Schryver, P.; De Gusseme, B.; De Muynck, W.; Boon, N.; Verstraete, W. Chemical and biological technologies for
hydrogen sulfide emission control in sewer systems: A review. Water Res. 2008, 42, 1–12. [CrossRef] [PubMed]
65. Sivret, E.C.; Wang, B.; Parcsi, G.; Stuetz, R.M.; Sivret, E.C.; Wang, B.; Parcsi, G.; Stuetz, R.M. Prioritisation of odorants emitted
from sewers using odour activity values. Water Res. 2016, 88, 308–321. [CrossRef]
66. Hwang, Y.; Matsuo, T.; Hanaki, K.; Suzuki, N. Identification and quantification of sulfur and nitrogen containing odorous
compounds in wastewater. Water Res. 1995, 29, 711–718. [CrossRef]
67. Wang, B.; Sivret, E.C.; Parcsi, G.; Le, N.M.; Kenny, S.; Bustamante, H.; Stuetz, R.M. Reduced sulfur compounds in the atmosphere
of sewer networks in Australia: Geographic (and seasonal) variations. Water Sci. Technol. 2013, 69, 1167–1173. [CrossRef]
Energies 2022, 15, 4015 15 of 15

68. Devai, I.; DeLaune, R.D. Emission of Reduced Malodorous Sulfur Gases from Wastewater Treatment Plants. Water Environ. Res.
1999, 71, 203–208. [CrossRef]
69. Chan, A.A.; Han, J. Odorous wastewater emissions. Vatten 2006, 62, 227–236. Available online: https://www.semanticscholar.
org/paper/Odorous-wastewater-emissions-Andersson-Chan-Hanaeus/56e586d26ed89a467638a6ee7f969cbcf983b79a (accessed
on 24 April 2022).
70. Lasaridi, K.; Katsabanis, G.; Kyriacou, A.; Maggos, T.; Manios, T.; Fountoulakis, M.; Stentiford, E.I. Assessing odour nuisance
from wastewater treatment and composting facilities in Greece. Waste Manag. Res. 2010, 28, 977–984. [CrossRef]
71. Pérez, A.; Manjón, C.; Martínez, J.V.; Juárez-Galan, J.M.; Barillon, B.; Bouchy, L. Odours in sewer networks: Nuisance assessment.
Water Sci. Technol. 2013, 67, 543–548. [CrossRef]
72. Gostelow, P.; Parsons, S.A.; Stuetz, R.M. Odour measurements for sewage treatment works. Water Res. 2001, 35,
579–597. [CrossRef]
73. Thistlethwayte, D.K.B.; Goleb, E.E. The composition of sewer air. In Proceedings of the 6th International Conference on Water
Pollution Research, Jerusalem, Israel, 18–23 June 1972; pp. 281–289.
74. Van Gemert, L.J. Odour Thresholds: Compilations of Odour Threshold Values in Air, Water and Other Media; Oliemans Punter & Partners
BV: Zeist, The Netherlands, 2003.
75. Sutherland-Stacey, L.; Corrie, S.; Neethling, A.; Johnson, I.; Gutierrez, O.; Dexter, R.; Yuan, Z.; Keller, J.; Hamilton, G. Continuous
measurement of dissolved sulfide in sewer systems. Water Sci. Technol. 2008, 57, 375–381. [CrossRef] [PubMed]
76. Tchobanoglous, G.; Franklin, L.M.; Burton, E.C.; Stensel, H.D. Wastewater Engineering Treatment and Reuse, 4th ed.; McGraw-Hill:
New York, NY, USA, 2011.
77. Eaton, A.D.; Franson, M.A.H.; Clesceri, L.S.; Rice, E.W.; Greenberg, A.E.; American Public Health Association; American Water
Works Association; Water Environment Federation. Standard Methods for the Examination of Water & Wastewater; American Public
Health Association: Washington, DC, USA, 2005; ISBN 0875530478.
78. Shammay, A.; Evason, I.E.J.; Stuetz, R.M. Prioritisation of odorants emitted from sewers using odour activity values. J. Environ.
Manag. 2019, 249, 308–321. [CrossRef]

You might also like